Next Article in Journal
Characterisation of the First Archaeal Mannonate Dehydratase from Thermoplasma acidophilum and Its Potential Role in the Catabolism of D-Mannose
Next Article in Special Issue
Facet-Dependent Reactivity of Fe2O3/CeO2 Nanocomposites: Effect of Ceria Morphology on CO Oxidation
Previous Article in Journal
Overcoming Water Insolubility in Flow: Enantioselective Hydrolysis of Naproxen Ester
Previous Article in Special Issue
The Support Effects on the Direct Conversion of Syngas to Higher Alcohol Synthesis over Copper-Based Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ceria Nanoparticles’ Morphological Effects on the N2O Decomposition Performance of Co3O4/CeO2 Mixed Oxides

by
Maria Lykaki
1,
Eleni Papista
2,
Nikolaos Kaklidis
2,
Sόnia A. C. Carabineiro
3 and
Michalis Konsolakis
1,*
1
School of Production Engineering and Management, Technical University of Crete, 73100 Chania, Greece
2
Department of Mechanical Engineering, University of Western Macedonia, GR-50100 Kozani, Greece
3
Laboratório de Catálise e Materiais (LCM), Laboratório Associado LSRE-LCM, Faculdade de Engenharia, Universidade do Porto, 4200-465 Porto, Portugal
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(3), 233; https://doi.org/10.3390/catal9030233
Submission received: 15 January 2019 / Revised: 15 February 2019 / Accepted: 18 February 2019 / Published: 3 March 2019

Abstract

:
Ceria-based oxides have been widely explored recently in the direct decomposition of N2O (deN2O) due to their unique redox/surface properties and lower cost as compared to noble metal-based catalysts. Cobalt oxide dispersed on ceria is among the most active mixed oxides with its efficiency strongly affected by counterpart features, such as particle size and morphology. In this work, the morphological effect of ceria nanostructures (nanorods (ΝR), nanocubes (NC), nanopolyhedra (NP)) on the solid-state properties and the deN2O performance of the Co3O4/CeO2 binary system is investigated. Several characterization methods involving N2 adsorption at −196 °C, X-ray diffraction (XRD), temperature programmed reduction (TPR), X-ray photoelectron spectroscopy (XPS) and transmission electron microscopy (ΤΕΜ) were carried out to disclose structure–property relationships. The results revealed the importance of support morphology on the physicochemical properties and the N2O conversion performance of bare ceria samples, following the order nanorods (NR) > nanopolyhedra (NP) > nanocubes (NC). More importantly, Co3O4 impregnation to different carriers towards the formation of Co3O4/CeO2 mixed oxides greatly enhanced the deN2O performance as compared to bare ceria samples, without, however, affecting the conversion sequence, implying the pivotal role of ceria support. The Co3O4/CeO2 sample with the rod-like morphology exhibited the best deN2O performance (100% N2O conversion at 500 °C) due to its abundance in Co2+ active sites and Ce3+ species in conjunction to its improved reducibility, oxygen kinetics and surface area.

Graphical Abstract

1. Introduction

Nitrous oxide (N2O) is one of the most significant greenhouse gases contributing to the depletion of the ozone layer. N2O has a much higher global warming potential (GWP) compared to CO2 (310 times higher) and a long atmospheric lifetime (114 years). The emissions of N2O are derived by both natural and anthropogenic sources. The main anthropogenic sources for N2O emissions involve agriculture (use of fertilizers), chemical industry (adipic and nitric acid production), the combustion of fossil fuels, as well as biomass burning, etc. [1,2,3,4].
The abatement of N2O emissions is of paramount importance and the direct catalytic decomposition of nitrous oxide to molecular nitrogen and oxygen (deN2O process) is considered to be a highly efficient remediation method. Thus far, several catalytic systems, such as supported noble metals [5,6,7], perovskites [8,9,10], hexaaluminates [11,12,13,14], spinels [15,16,17,18], zeolites [19,20,21,22] and mixed oxides [23,24,25,26,27], have been used for N2O decomposition. Although noble metals exhibit satisfactory activity for the deN2O process, their high cost and the deterioration of their catalytic efficiency from gases present in the exhaust gas stream (e.g., O2) act as inhibiting factors for practical applications [1,28]. Hence, research efforts have focused on the development of noble metal-free mixed oxides of high activity, stability and low cost, as recently reviewed [1].
Among the different transition metal oxides, cobalt spinel shows unique physicochemical characteristics, such as thermal stability and high reducibility, making it an excellent candidate for the deN2O process [15,23,29,30]. However, the high cost of cobalt renders mandatory its dispersion to high surface area supports like ceria, magnesia, etc. [31,32]. Among the various supports investigated, ceria exhibits unique redox properties associated with its high oxygen storage capacity (OSC), rendering this material highly effective in many catalytic processes [23,33,34,35]. Furthermore, the synergistic effects induced by strong metal–ceria interactions, in nanoscale, can modify the surface chemistry of the materials through geometric or/and electronic perturbations, leading to improved redox properties and catalytic activity [36,37,38,39,40].
However, the catalytic efficiency of transition metal oxides, involving ceria-based mixed oxides, can be considerably affected by the different counterpart characteristics, such as particle size and morphology. In this regard, engineering the particle size and shape (e.g., nanorods and nanocubes) through the employment of advanced nano-synthesis paths has lately received particular attention [33,41,42,43]. Interestingly, the support morphology greatly affects the redox properties, oxygen mobility and, subsequently, the catalytic activity of the mixed oxides. For instance, Lin et al. [44] prepared Co3O4/CeO2 catalysts with three different support morphologies, namely polyhedra, nanorods and hexagonal shapes, with polyhedra exhibiting the highest catalytic activity for ammonia synthesis. In a similar manner, by tailoring the support morphology, CuO/CeO2 nanoshaped materials of enhanced reducibility and deN2O performance can be obtained [45]. Andrade-Martínez et al. [46] investigated the catalytic reduction of N2O over CuO/SiO2 catalysts, revealing the key role of the spherical ordered mesoporous support, along with its functionalization through copper addition, on the improved catalytic activity and stability, making this material comparable to noble metal-reported systems. Different support morphologies (rods, plates and cubes) have also been employed for the low temperature CH2Br2 oxidation revealing the superiority of cobalt-ceria nanorods in the catalytic performance [47]. Moreover, cobalt oxide supported on ceria of different morphology (nanoparticles, nanorods and nanocubes) has been investigated for the catalytic oxidation of toluene with the nanoparticles exhibiting the highest catalytic activity due to the synergism at the interface between the two oxide phases, which leads to an improved reducibility [48]. Very recently, the influence of support morphology (nanorods, nanocubes and nanopolyhedra) on the surface and structural properties of CuO/CeO2 mixed oxide has been thoroughly explored through both in situ and ex situ characterization techniques. The results disclosed the significance of the ceria morphology on the reducibility and oxygen kinetics, revealing the order nanorods > nanopolyhedra > nanocubes [49].
In this work, ceria structures of various morphologies (nanopolyhedra, nanorods and nanocubes) were hydrothermally prepared, and then cobalt was impregnated into the above ceria supports. The purpose of this work was to explore the impact of support morphology on the surface chemistry and the deN2O performance of Co3O4/CeO2 mixed oxides. The results clearly revealed that support morphology can exert a profound influence on the N2O decomposition, paving the way toward the rational design of highly efficient deN2O catalysts.

2. Results and Discussion

2.1. Textural/Structural Analysis (BET and XRD)

The main textural and structural characteristics of bare ceria samples and Co3O4/CeO2 mixed oxides (hereinafter denoted as Co/CeO2) are summarized in Table 1. According to the BET surface area, the following order is acquired: CeO2-NP (88 m2 g−1) > CeO2-NR (79 m2 g−1) > CeO2-NC (37 m2 g−1). The addition of cobalt into CeO2 decreases the surface area, without, however, significantly affecting the order obtained for bare ceria samples. The Co/CeO2-NR sample exhibits the highest value (72 m2 g−1) succeeded by Co/CeO2-NP (71 m2 g−1) and Co/CeO2-NC (28 m2 g−1). Regarding the average pore diameter and pore volume, they both decreased upon the addition of Co to ceria nanorods and nanocubes. However, concerning ceria nanopolyhedra, the addition of cobalt leads to a small increase in the average pore diameter, whereas the pore volume is not significantly affected (Table 1).
Figure 1a shows the Barret-Joyner-Halenda (BJH) desorption pore size distributions (PSD) of bare ceria and Co/CeO2 catalysts. According to the pore size distribution, all the samples have their maxima at a pore diameter more than 3 nm, designating the presence of mesopores [50]. It is obvious that bare ceria samples with the nanocube (CeO2-NC) and nanorod morphology (CeO2-NR) exhibit similar particle size distributions, whereas in ceria nanopolyhedra (CeO2-NP), a narrower PSD is observed. Noteworthily, PSD remains practically unaffected by the addition of cobalt in all cases. As it can be observed in Figure 1b which shows the adsorption–desorption isotherms, all samples demonstrate type IV isotherms with a hysteresis loop at a relative pressure > 0.5, further corroborating the mesoporous structure of the materials [51,52].
The XRD patterns of the samples are shown in Figure 2. The main peaks can be indexed to (111), (200), (220), (311), (222), (400), (331), (420), (422), (511) and (440) planes which are attributed to ceria face-centered cubic fluorite structure (Fm3m symmetry, no. 225) [53,54]. There are three small peaks at 2θ values of approx. 36, 44 and 64o which are typical of Co3O4 [33]. These three diffraction peaks correspond to the (311), (400) and (440) planes of Co3O4, respectively. The average crystallite diameter of the oxide phases (CeO2 and Co3O4) was assessed by an XRD analysis by means of the Williamson–Hall plot (Table 1). The CeO2 crystallite size measurements showed 24, 14 and 11 nm for Co/CeO2-NC, Co/CeO2-NR and Co/CeO2-NP, respectively. As it is obvious from Table 1, there is a small decrease in the ceria crystallite size for nanocubes and nanorods, whereas no changes were observed for nanopolyhedra, indicating that the structural characteristics of ceria supports do not get significantly affected upon cobalt addition, as it will be further corroborated by a TEM analysis (see below). In a similar manner, the BET analysis (Table 1) indicates no significant modifications on the pore characteristics of ceria nanopolyhedra upon cobalt addition, which could be ascribed to their irregular morphology. It should be also noted that the samples with nanocube morphology exhibit the smallest BET surface area and the largest CeO2 and Co3O4 crystallite sizes in comparison to nanorods and nanopolyhedra. As for the crystallite size of cobalt oxide phase, the following sequence was obtained: Co/CeO2-NC (19 nm) > Co/CeO2-NR (16 nm) > Co/CeO2-NP (15 nm), which perfectly matches the order obtained for CeO2.

2.2. Morphological Characterization (TEM)

Transmission electron microscopy (TEM) has been applied so as to examine the morphological differences among the materials. Figure 3a–c shows the TEM images of ceria supports. The CeO2-NR sample (Figure 3a) exhibits a rod-shaped morphology with the length varying between 25 and 200 nm. Figure 3b and c demonstrates mainly irregular-shaped nanopolyhedra and cubes, respectively. Figure 3d–f illustrates the images derived by TEM analysis for the Co/CeO2 mixed oxides. Evidently, the morphology is not affected by cobalt addition to the ceria carrier.

2.3. Redox Properties (H2-Temperature Programmed Reduction (TPR))

H2-TPR experiments took place to investigate the ceria shape effect on the redox properties of as-prepared samples. Figure 4a shows the TPR profiles of bare ceria samples, consisting of two wide-ranging peaks which are centred at 526–551 °C and 789–813 °C. These peaks are attributed to ceria surface oxygen (Os) and bulk oxygen (Ob) reduction, respectively [33,49,55]. In Table 2, the hydrogen consumption corresponding to surface oxygen as well as to bulk oxygen reduction is presented. Based on the ratio of surface-to-bulk oxygen (Os/Ob), the following order was acquired: CeO2-NR (1.13) > CeO2-NP (0.94) > CeO2-NC (0.71). This indicates the superior reducibility of the rod-shaped sample as it exhibits the highest amount of loosely bound oxygen species. The latter is expected to notably affect the deN2O process, where the desorption of adsorbed oxygen species mainly determines the reaction rate (vide infra).
The reduction profiles of the Co/CeO2 samples as well as the one of a Co3O4 reference are shown in Figure 4b. Table 2 summarizes the main TPR peaks along with the hydrogen consumption (mmol H2 g−1). Pure Co3O4 shows two reduction peaks (a and b) in much lower temperatures than those of bare ceria samples, namely 305 °C and 415 °C. They are ascribed to the stepwise reduction of Co3O4 → CoO → Co, respectively [44,56,57,58].
On the other hand, Co/CeO2 samples exhibit two main peaks at the temperature range of 318–335 °C (peak a) and 388–405 °C (peak b), ascribed to the reduction of Co3+ to Co2+ and Co2+ to Co0, respectively [33,59,60]. Obviously, the cobalt addition facilitates the reduction of ceria surface oxygen, shifting the peaks centered at 526–551 °C to a lower temperature (comparison of Figure 4a,b). They also exhibit a broad peak above 800 °C, attributed to the ceria subsurface oxygen reduction, while the capping oxygen reduction overlaps with the reduction of CoO [33,56,61]. Apparently, the reduction of the mixed oxides takes place in lower temperatures compared to the bare ceria samples, demonstrating the beneficial effect of cobalt on the surface oxygen reduction of ceria. In fact, the interaction between the two oxide phases could be considered responsible for the improved reducibility and oxygen mobility, as thoroughly discussed in previous studies [48,49,62]. According to the consumption of hydrogen in the low-temperature range (Table 2), which could be related to the cobalt species reduction along with the ceria surface oxygen reduction, the Co/CeO2-NP and Co/CeO2-NR samples exhibit a similar H2 uptake (about 2.40 mmol H2 g−1) while the sample of nanocube morphology exhibits a much lower value (2.05 mmol H2 g−1). This trend is well-matched to the catalytic results (vide infra), revealing the key role of redox ability on the deN2O process.
Moreover, the Co/CeO2-NR sample exhibits the lowest reduction temperature (peak at 318 °C) in comparison with the other samples (peak ca. 335 °C), indicating the facilitation of Co3+ species reduction over ceria nanorods. Noteworthily, the theoretical amount of hydrogen for the complete reduction of Co3O4 to Co (approx. 1.76 mmol H2 g−1, based on a 7.8 wt. % nominal loading of Co) is always surpassed by the hydrogen amount required for the reduction of Co/CeO2 samples (Table 2). The latter reveals the facilitation of ceria capping oxygen reduction in the presence of cobalt, further corroborating the above findings.

2.4. Surface Analysis (X-ray Photoelectron Spectroscopy (XPS))

An XPS analysis was performed in order to investigate the effect of ceria morphology on the elemental chemical states and surface composition of Co/CeO2 mixed oxides. Figure 5a shows the Ce3d XPS spectra of ceria nanoparticles of different morphology and the corresponding Co/CeO2 samples, which can be deconvoluted into eight components [63,64,65], with the assignment of the characteristic peaks having been thoroughly described in our previous work [49]. In brief, the three pairs of peaks labeled as u, v; u", v"; and u‴, v‴ are ascribed to Ce4+, whereas the residual u’ and v’ peaks are ascribed to Ce3+ species.
The corresponding O 1s spectra of the samples are depicted in Figure 5b. The low binding energy peak at 529.3 eV is attributed to the lattice oxygen (OI) of Co3O4 and CeO2 phases, and the high binding energy peak at 531.3 eV corresponds to the chemisorbed oxygen (OII) such as adsorbed oxygen (O/O22−) and water, carbonate as well as hydroxyl species [23,56].
The proportion of Ce3+ (%) as well as the OII/OI ratio for all samples is summarized in Table 3. Bare ceria supports exhibit a similar amount of Ce3+ ranging from 23.3 to 25.3%. Regarding, the cobalt-ceria samples, the population of Ce3+ species is slightly higher, varying between 26.1 and 28.5%. In particular, the Co/CeO2-NR sample exhibits the highest amount (28.5%) followed by Co/CeO2-NP (26.7%) and Co/CeO2-NC (26.1%), indicating the abundance of the nanorod samples in oxygen vacancies. Interestingly, the relative ratio of adsorbed to lattice oxygen (OII/OI) and the Ce3+ (%) follow the same order, namely, Co/CeO2-NR (0.60) > Co/CeO2-NP (0.53) > Co/CeO2-NC (0.51), perfectly matched to the order obtained for the catalytic performance, as it will be discussed in the sequence. It should be also noted that Co addition to CeO2-NR enhances both the population of reduced Ce3+ species and the OII/OI ratio, revealing the synergistic interactions between cerium and cobalt oxides toward the formation of highly reducible composites, in agreement with the TPR results.
Figure 6 depicts the Co 2p XPS spectra of Co/CeO2 samples along with the spectrum obtained for the Co3O4 reference sample for comparison purposes. The samples exhibit two major peaks of Co2p3/2 (780 eV) and Co2p1/2 (795 eV). According to peaks’ positions and shapes, the structure of the cobalt spinel is formed [23,66,67]. The Co2+/Co3+ ratio of Co/CeO2 samples derived by the deconvolution of the Co2p1/2 and Co2p3/2 peaks is included in Table 3. The nanorod sample, which offers the best deN2O performance (vide infra), exhibits the highest Co2+/Co3+ ratio (1.32), followed by nanocubes (1.06) and nanopolyhedra (0.94). In view of this fact, it has been reported that samples with a high Co2+/Co3+ ratio exhibit better deN2O performance [3,20,22,43,59], further corroborating the present findings.

2.5. Catalytic Evaluation Studies

The impact of ceria morphology on the catalytic decomposition of N2O under oxygen deficient and oxygen excess conditions was next examined. Figure 7a,b shows the N2O conversion profiles as a temperature function for bare ceria as well as Co/CeO2 samples in the absence and presence of oxygen, respectively. The Co/CeO2-NR sample exhibits the best catalytic performance, both in the absence and presence of oxygen in the gas stream. Apparently, the addition of cobalt in the ceria lattice enormously enhances the catalytic efficiency without, however, affecting the catalytic order of bare ceria samples, suggesting the pivotal role of ceria morphology on the deN2O performance. In terms of the half-conversion temperature (T50), the following order is obtained for the mixed oxides in the absence of oxygen: Co/CeO2-NR (449 °C) > Co/CeO2-NP (458 °C) > Co/CeO2-NC (464 °C). The same trend is observed in the presence of oxygen as well, although in slightly higher temperatures, due to its competitive sorption on the catalyst surface. In this point, it should be noted that the un-catalyzed reaction shows nearly zero N2O conversion in the temperature range investigated, as previously reported [29,46,68].
The above findings can be well-interpreted by taking into account a redox-type mechanism for the decomposition of N2O over cobalt spinel oxides [4,23,24,30,59,69,70,71,72,73]:
Co2+ + N2O → Co3+–O + N2
Co3+–O + N2O → Co3+–O2 + N2
Co3+–O2 → Co2+ + O2
In this mechanistic sequence, N2O is initially chemisorbed on the Co2+ sites (Equation (1)) which are considered as the active centres for initiating the N2O dissociative adsorption. Then, the regeneration of the active sites is taking place through the Co3+/Co2+ redox cycle, involving the combination of O into O2 (Equation (2)) and the desorption of molecular oxygen (Equation (3)), which finally leads to the regeneration of those sites [69].
However, in the case of Co3O4/CeO2 mixed oxides, the excellent redox characteristics of ceria, such as oxygen storage capacity and oxygen mobility, can be further accounted for the regeneration of active sites through the following steps [69]:
Co3+–O + Ce3+–Ovac → Co2+ + Ce4+–O
2Ce4+–O ↔ Ce4+–O22−–Ce4+
Ce4+–O + N2O → Ce3+–Ovac + N2 + O2
Based on the above mechanistic scheme, the superiority of the Co/CeO2 sample with a rod-like morphology can receive a consistent explanation. More specifically, nanorod-shaped ceria with (110) and (100) reactive planes exhibit enhanced oxygen kinetics and reducibility as it has the highest population of loosely bound oxygen species (Table 2), which is a decisive factor in terms of deN2O activity. In other words, the high amount of weakly bound oxygen species present in the Co3O4/CeO2 samples of rod-like morphology, linked directly to oxygen vacancy formation and oxygen mobility, could be considered responsible for the formation and the consequent regeneration of active sites. In this regard, a perfect interrelation between the catalytic performance (in terms of the half-conversion temperature, T50) and the redox properties (in terms of the ratio of surface oxygen to bulk oxygen, Os/Ob) is disclosed, as illustrated in Figure 8. This clearly justifies the key role of redox properties on the deN2O process. In a similar manner, Liu et al. [28] have pointed out that the synergistic interaction between the two oxide phases in a CuO–CeO2 mixed oxide enhances the reducibility and consequently the deN2O efficiency as the surface-adsorbed oxygen species is easily desorbed and the active sites’ regeneration is enabled.
More interestingly, the deN2O performance of CeO2 as well as the Co3O4/CeO2 samples totally coincides, indicating the significance of the ceria carrier. However, the superiority of the mixed oxides in comparison to the bare ceria samples is evident, reflecting the synergistic interactions between cobalt and cerium oxides. The latter is manifested by the improved redox properties (in terms of H2 consumption and TPR onset temperature) of Co3O4/CeO2 mixed oxides as compared to bare ceria (Table 2). In a similar manner, the incorporation of cobalt into the ceria lattice increases both the amount of the adsorbed oxygen species (O/O22−) and Ce3+ (Table 3), related with the surface oxygen reduction and the abundance in oxygen vacancies (Ovac).
Moreover, ceria nanorods facilitate the reduction of Co3+ to Co2+ active sites (Table 3), further contributing to the superior catalytic performance of the Co/CeO2-NR sample. Along the same line, it has been recently reported that ceria nanorods stabilize the partial oxidation state of Co in CoOX/CeO2 catalysts via the facilitation of oxygen transfer at the metal-support interface [74]. It should be, therefore, deduced that ceria nanorods with the exposed (110) and (100) facets show the highest surface-to-bulk oxygen ratio resulting in improved reducibility and oxygen kinetics while exhibiting the highest amount of weakly bound oxygen species which is a decisive factor in the deN2O process. Upon cobalt addition, the nanorod sample exhibits in addition the highest population in Ce3+/Co2+ redox pairs, indicative of abundant oxygen vacancies, which, along with its enhanced reducibility, leads to a superior deN2O performance.
In this point, the enhanced textural characteristics (BET area and pore volume) of Co/CeO2-NR as compared to Co/CeO2-NC should be also mentioned, which could be further accounted for its enhanced deN2O performance. Thus, by taking into account the specific activity normalized per unit of surface area (nmol m−2 s−1) instead of mass unit (nmol g−1 s−1), an inferior performance is observed for Co/CeO2-NR compared to Co/CeO2-NC (Table 4). On the other hand, Co3O4/CeO2-NR exhibits a superior deN2O performance (both in terms of conversion and specific activity) as compared to Co3O4/CeO2-NP despite their similar structural (crystallite size) and textural (surface area) properties (Table 1). The latter clearly reveals the importance of exposed facets and redox properties on the deN2O process, as it has been similarly reported by Zabilskiy et al. [45] for CuO/CeO2 nanostructures of different morphology. Therefore, on the basis of the present findings, it can be deduced that the enhanced N2O conversion performance of Co3O4/CeO2-NR mixed oxides could be attributed to a compromise between redox and textural characteristics.

3. Materials and Methods

3.1. Materials Synthesis

In the present work, the chemicals that were used were of analytical reagent grade. Ce(NO3)3·6H2O (Fluka, Bucharest, Romania, purity ≥99.0%) and Co(NO3)2·6H2O (Sigma-Aldrich, Taufkirchen, Germany, purity ≥98%) were employed as precursor compounds for the preparation of bare ceria as well as Co/CeO2 mixed oxides. Also, NaOH (Sigma-Aldrich, Taufkirchen, Germany, purity ≥98%) and ethanol (ACROS Organics, Geel, Belgium, purity 99.8%,) were used during materials synthesis. Initially, the hydrothermal method was applied for the preparation of bare ceria nanoparticles, as described in detail in our previous work [49]. In brief, ceria nanorods (CeO2-NR) were synthesized by dissolving NaOH (36.7 M) in double deionized water and then adding an appropriate amount of an aqueous solution of Ce(NO3)3·6H2O (0.13 M) under vigorous stirring. Next, the transfer of the final slurry into a Teflon bottle and its aging at 90 °C for 24 h occurred. For the synthesis of ceria nanopolyhedra (CeO2-NP), a similar procedure was followed, employing, however, a lower amount of NaOH (6 M). In order to synthesize ceria nanocubes (CeO2-NC), the same procedure as the one described above for the synthesis of ceria nanorods was followed, with the obtained slurry to be aged at 180 °C instead of 90 °C. In all cases, centrifugation was used for the recovery of the solid products that were thoroughly washed with double deionized water until a neutral pH was reached and finally washed with ethanol so as to avoid the nanoparticles’ agglomeration. Afterwards, drying of the precipitate at 90 °C for 12 h followed by calcination at 500 °C for 2 h under air flow (heating ramp 5 °C/min) was carried out.
The Co/CeO2-NX catalysts where NX stands for NP: nanopolyhedra, NR: nanorods and NC: nanocubes were prepared by wet impregnation, employing an aqueous solution of Co(NO3)2·6H2O, in order to achieve an atomic ratio Co/(Co+Ce) of 0.2 which corresponds to 7.8 wt. % of Co loading. Heating under stirring of the obtained suspensions until complete water evaporation occurred, followed by drying at 90 °C for 12 h and final calcination at 500 °C for 2 h under air flow (heating ramp 5 °C/min).

3.2. Materials Characterization

The porosity of the materials was evaluated by the N2-adsorption isotherms at −196 °C, using an ASAP 2010 (Micromeritics, Norcross, GA, USA) apparatus (from ReQuimTe Analyses Laboratory, Universidade Nova de Lisboa, Lisboa, Portugal). The samples were previously degassed at 300 °C for 6 h. The specific surface area was calculated by the Brunauer–Emmett–Teller (BET) equation [75].
Structural characterization was carried out by means of X-ray diffraction (XRD) in a PAN’alytical X’Pert MPD equipped with a X’Celerator detector and secondary monochromator (Cu Kα λ = 0.154 nm, 50 kV, 40 mA; data recorded at a 0.017º step size, 100 s/step) in the University of Trás-os-Montes e Alto Douro. The collected spectra were analyzed by Rietveld refinement using PowderCell software, allowing the determination of crystallite sizes by means of the Williamson–Hall plot.
The redox properties were assessed by Temperature Programmed Reduction (TPR) experiments in an AMI-200 Catalyst Characterization Instrument (Altamira Instruments, Pittsburgh, PA, USA), employing H2 as a reducing agent. In a typical H2-TPR experiment, 50 mg of the sample (grain size 180–355 μm) was heated up to 1100 °C (10 °C/min) under H2 flow (1.5 cm3) balanced with He (29 cm3). The amount of H2 consumed (mmol g−1) was calculated by taking into account the integrated area of TPR peaks, calibrated against a known amount of CuO standard sample [76,77].
The surface composition and the chemical state of each element were determined by X-ray photoelectron spectroscopy (XPS) analyses, performed on a VG Scientific ESCALAB 200A spectrometer using Al Kα radiation (1486.6 eV) in CEMUP. The charge effect was corrected using the C1s’ peak as a reference (binding energy of 285 eV). The CASAXPS software was used for data analysis.
The samples were imaged by transmission electron microscopy (TEM). The analyses were performed on a Leo 906E apparatus (Austin, TX, USA), at 120 kV in the University of Trás-os-Montes e Alto Douro. The samples were prepared by ultrasonic dispersion, and a 400 mesh formvar/carbon copper grid (Agar Scientific, Essex, UK) was dipped into the solution for the TEM analysis.

3.3. Catalytic Performance Evaluation

The catalytic studies for the N2O decomposition took place in a quartz fixed-bed U-shaped reactor (0.8 cm i.d.) with 100 mg of catalyst loading (grain size 180–355 μm). The feed gas (1000 ppm N2O and 0 or 2 vol. % O2) total flow rate was 150 cm3/min which corresponds to a Gas Hour Space Velocity (GHSV) of 40,000 h−1. The analysis of the gases was performed by a gas chromatograph (SHIMADZU 14B). The apparatus is equipped with a TCD detector and two separation columns (Molecular Sieve 5A for O2, N2 measurements and Porapack QS for N2O measurement). Prior to the catalytic activity measurements, the materials under consideration were subjected to further processing under He flow (100 cm3/min) at 400 °C. In order to minimize the external and internal diffusion limitations, preliminary experiments concerning the influence of particle size and W/F ratio on deN2O catalytic performance were carried out. Based on these experiments, a catalyst particle size in the range of 180–355 μm was selected, in addition to a W/F ratio of 0.04 g s cm−3. The conversion of N2O (XN2O) was calculated from the difference in N2O concentration between the inlet and outlet gas streams, according to the equation
X N 2 O ( % )   =   [ N 2 O ] in [ N 2 O ] out [ N 2 O ] in × 100
The specific reaction rate (r, mol m−2 s−1) of the N2O decomposition was also estimated using the following formula:
r ( mol   m 2   s 1 )   =   X N 2 O × [ N 2 O ] in × F ( cm 3 min ) 100 × 60 ( s min ) × V m ( cm 3 mol ) × m cat ( g ) × S BET ( m 2 g )
where F and Vm are the total flow rate and gas molar volume, respectively, at standard ambient temperature and pressure conditions (298 K and 1 bar), mcat is the catalyst’s mass and SBET is the surface area.

4. Conclusions

In this work, three different ceria nanoshaped materials (nanorods, nanocubes and nanopolyhedra) were hydrothermally synthesized and used as supports for the cobalt oxide phase. Both single CeO2 and Co/CeO2 mixed oxides were catalytically assessed during the decomposition of N2O in the presence and absence of oxygen. For bare ceria samples, the following deN2O order was obtained: CeO2-NR (nanorods) > CeO2-NP (nanopolyhedra) > CeO2-NC (nanocubes). Most importantly, cobalt addition to the CeO2 carriers greatly enhances the N2O decomposition, not affecting at all the order obtained for the bare ceria supports and clearly reflecting the key role of support. The present results clearly reveal the key role of support morphology on the textural, structural and redox properties, reflected then on the catalytic performance of Co3O4/CeO2 mixed oxides. Among the different samples investigated, the cobalt-ceria nanorods (Co/CeO2-NR) exposing {100} and {110} facets showed the best deN2O performance, ascribed mainly to their abundance in Co2+ active sites in conjunction to their enhanced redox and textural properties.

Author Contributions

M.L. contributed to materials synthesis, results interpretation and paper writing; E.P. and N.K. contributed to catalytic evaluation studies; S.A.C.C. contributed to the materials characterization; M.K. contributed to the conception, design, results interpretation and writing of the paper; all authors contributed to the discussion and read and approved the final version of the manuscript.

Acknowledgments

The research work was supported by the Hellenic Foundation for Research and Innovation (HFRI) and the General Secretariat for Research and Technology (GSRT), under the HFRI PhD Fellowship grant (GA. no. 34252). This research has been cofinanced by the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH—CREATE—INNOVATE (project code: T1EDK-00094). This work was also financially supported by Associate Laboratory LSRE-LCM—UID/EQU/50020/2019—funded by national funds through FCT/MCTES (PIDDAC). S.A.C.C. acknowledges Fundação para a Ciência e a Tecnologia (Portugal) for Investigador FCT program (IF/01381/2013/CP1160/CT0007), with financing from the European Social Fund and the Human Potential Operational Program. We are grateful to Carlos Sá (CEMUP) for the assistance with the XPS measurements, Pedro Tavares (UTAD) for the TEM and XRD analyses and Nuno Costa (ReQuimTe) for the N2 adsorption results.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Konsolakis, M. Recent Advances on Nitrous Oxide (N2O) Decomposition over Non-Noble-Metal Oxide Catalysts: Catalytic Performance, Mechanistic Considerations and Surface Chemistry Aspects. ACS Catal. 2015, 5, 6397–6421. [Google Scholar] [CrossRef]
  2. Liu, Z.; He, F.; Ma, L.; Peng, S. Recent Advances in Catalytic Decomposition of N2O on Noble Metal and Metal Oxide Catalysts. Catal. Surv. Asia 2016, 20, 121–132. [Google Scholar] [CrossRef]
  3. Kim, M.J.; Lee, S.J.; Ryu, I.S.; Jeon, M.W.; Moon, S.H.; Roh, H.S.; Jeon, S.G. Catalytic decomposition of N2O over cobalt based spinel oxides: The role of additives. Mol. Catal. 2017, 442, 202–207. [Google Scholar] [CrossRef]
  4. Rutkowska, M. Catalytic Decomposition of N2O Using a New Generation of Functionalized Microporous and Mesoporous Inorganic Materials. Wiadomości Chemiczne 2015, 69, 297–315. [Google Scholar]
  5. Piumetti, M.; Hussain, M.; Fino, D.; Russo, N. Mesoporous silica supported Rh catalysts for high concentration N2O decomposition. Appl. Catal. B Environ. 2015, 165, 158–168. [Google Scholar] [CrossRef]
  6. Kim, S.S.; Lee, S.J.; Hong, S.C. Effect of CeO2 addition to Rh/Al2O3 catalyst on N2O decomposition. Chem. Eng. J. 2011, 169, 173–179. [Google Scholar] [CrossRef]
  7. Hussain, M.; Fino, D.; Russo, N. Development of modified KIT-6 and SBA-15-spherical supported Rh catalysts for N2O abatement: From powder to monolith supported catalysts. Chem. Eng. J. 2014, 238, 198–205. [Google Scholar] [CrossRef]
  8. Wu, Y.; Cordier, C.; Berrier, E.; Nuns, N.; Dujardin, C.; Granger, P. Surface reconstructions of LaCo1xFexO3 at high temperature during N2O decomposition in realistic exhaust gas composition: Impact on the catalytic properties. Appl. Catal. B Environ. 2013, 140–141, 151–163. [Google Scholar] [CrossRef]
  9. Ivanov, D.V.; Pinaeva, L.G.; Isupova, L.A.; Sadovskaya, E.M.; Prosvirin, I.P.; Gerasimov, E.Y.; Yakovleva, I.S. Effect of surface decoration with LaSrFeO4 on oxygen mobility and catalytic activity of La0.4Sr0.6FeO3−δ in high-temperature N2O decomposition, methane combustion and ammonia oxidation. Appl. Catal. A Gen. 2013, 457, 42–51. [Google Scholar] [CrossRef]
  10. Kumar, S.; Vinu, A.; Subrt, J.; Bakardjieva, S.; Rayalu, S.; Teraoka, Y.; Labhsetwar, N. Catalytic N2O decomposition on Pr0.8Ba0.2MnO3 type perovskite catalyst for industrial emission control. Catal. Today 2012, 198, 125–132. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Wang, X.; Zhu, Y.; Zhang, T. Stabilization mechanism and crystallographic sites of Ru in Fe-promoted barium hexaaluminate under high-temperature condition for N2O decomposition. Appl. Catal. B Environ. 2013, 129, 382–393. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Wang, X.; Zhu, Y.; Hou, B.; Yang, X.; Liu, X.; Wang, J.; Li, J.; Zhang, T. Characterization of Fe substitution into La-hexaaluminate systems and the effect on N2O catalytic decomposition. J. Phys. Chem. C 2014, 118, 1999–2010. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Wang, X.; Zhu, Y.; Liu, X.; Zhang, T. Thermal evolution crystal structure and Fe crystallographic sites in LaFexAl12-xO19 hexaaluminates. J. Phys. Chem. C 2014, 118, 10792–10804. [Google Scholar] [CrossRef]
  14. Pérez-Ramírez, J.; Santiago, M. Metal-substituted hexaaluminates for high-temperature N2O abatement. Chem. Commun. 2007, 2, 619–621. [Google Scholar] [CrossRef] [PubMed]
  15. Stelmachowski, P.; Maniak, G.; Kaczmarczyk, J.; Zasada, F.; Piskorz, W.; Kotarba, A.; Sojka, Z. Mg and Al substituted cobalt spinels as catalysts for low temperature deN2O-Evidence for octahedral cobalt active sites. Appl. Catal. B Environ. 2014, 146, 105–111. [Google Scholar] [CrossRef]
  16. Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Duch, J.; Ćmil, K.; Kotarba, A.; Sojka, Z. Insights into the twofold role of Cs doping on deN2O activity of cobalt spinel catalyst-towards rational optimization of the precursor and loading. Appl. Catal. B Environ. 2015, 168–169, 509–514. [Google Scholar] [CrossRef]
  17. Zasada, F.; Piskorz, W.; Janas, J.; Gryboś, J.; Indyka, P.; Sojka, Z. Reactive Oxygen Species on the (100) Facet of Cobalt Spinel Nanocatalyst and their Relevance in 16O2/18O2 Isotopic Exchange, deN2O, and deCH4 Processes-A Theoretical and Experimental Account. ACS Catal. 2015, 5, 6879–6892. [Google Scholar] [CrossRef]
  18. Amrousse, R.; Chang, P.-J.; Choklati, A.; Friche, A.; Rai, M.; Bachar, A.; Follet-Houttemane, C.; Hori, K. Catalytic decomposition of N2O over Ni and Mg-magnetite catalysts. Catal. Sci. Technol. 2013, 3, 2288. [Google Scholar] [CrossRef]
  19. Zou, W.; Xie, P.; Hua, W.; Wang, Y.; Kong, D.; Yue, Y.; Ma, Z.; Yang, W.; Gao, Z. Catalytic decomposition of N2O over Cu-ZSM-5 nanosheets. J. Mol. Catal. A Chem. 2014, 394, 83–88. [Google Scholar] [CrossRef]
  20. Zhang, X.; Shen, Q.; He, C.; Ma, C.; Cheng, J.; Liu, Z.; Hao, Z. Decomposition of nitrous oxide over Co-zeolite catalysts: role of zeolite structure and active site. Catal. Sci. Technol. 2012, 2, 1249–1258. [Google Scholar] [CrossRef]
  21. Rutkowska, M.; Piwowarska, Z.; Micek, E.; Chmielarz, L. Hierarchical Fe-, Cu- and Co-Beta zeolites obtained by mesotemplate-free method. Part I: Synthesis and catalytic activity in N2O decomposition. Microporous Mesoporous Mater. 2015, 209, 54–65. [Google Scholar] [CrossRef]
  22. Xie, P.; Luo, Y.; Ma, Z.; Wang, L.; Huang, C.; Yue, Y.; Hua, W.; Gao, Z. CoZSM-11 catalysts for N2O decomposition: Effect of preparation methods and nature of active sites. Appl. Catal. B Environ. 2015, 170–171, 34–42. [Google Scholar] [CrossRef]
  23. Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Indyka, P.; Sojka, Z.; Guillén-Hurtado, N.; Rico-Pérez, V.; Bueno-López, A.; Kotarba, A. Strong dispersion effect of cobalt spinel active phase spread over ceria for catalytic N2O decomposition: The role of the interface periphery. Appl. Catal. B Environ. 2016, 180, 622–629. [Google Scholar] [CrossRef]
  24. Chromčáková, Ž.; Obalová, L.; Kovanda, F.; Legut, D.; Titov, A.; Ritz, M.; Fridrichová, D.; Michalik, S.; Kuśtrowski, P.; Jirátová, K. Effect of precursor synthesis on catalytic activity of Co3O4 in N2O decomposition. Catal. Today 2015, 257, 18–25. [Google Scholar] [CrossRef]
  25. Franken, T.; Palkovits, R. Investigation of potassium doped mixed spinels CuxCo3-xO4 as catalysts for an efficient N2O decomposition in real reaction conditions. Appl. Catal. B Environ. 2015, 176–177, 298–305. [Google Scholar] [CrossRef]
  26. Zabilskiy, M.; Erjavec, B.; Djinović, P.; Pintar, A. Ordered mesoporous CuO-CeO2 mixed oxides as an effective catalyst for N2O decomposition. Chem. Eng. J. 2014, 254, 153–162. [Google Scholar] [CrossRef]
  27. Xue, L.; He, H.; Liu, C.; Zhang, C.; Zhang, B. Promotion Effects and Mechanism of Alkali Metals and Alkaline Earth Metals on Cobalt - Cerium Composite Oxide Catalysts for N2O Decomposition. Environ. Sci. Technol. 2009, 43, 890–895. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, Z.; He, C.; Chen, B.; Liu, H. CuO-CeO2 mixed oxide catalyst for the catalytic decomposition of N2O in the presence of oxygen. Catal. Today 2017, 297, 78–83. [Google Scholar] [CrossRef]
  29. Russo, N.; Fino, D.; Saracco, G.; Specchia, V. N2O catalytic decomposition over various spinel-type oxides. Catal. Today 2007, 119, 228–232. [Google Scholar] [CrossRef]
  30. Xue, L.; Zhang, C.; He, H.; Teraoka, Y. Catalytic decomposition of N2O over CeO2 promoted Co3O4 spinel catalyst. Appl. Catal. B Environ. 2007, 75, 167–174. [Google Scholar] [CrossRef]
  31. Yang, J.; Lukashuk, L.; Akbarzadeh, J.; Stöger-Pollach, M.; Peterlik, H.; Föttinger, K.; Rupprechter, G.; Schubert, U. Different Synthesis Protocols for Co3O4-CeO2 Catalysts-Part 1: Influence on the Morphology on the Nanoscale. Chem. Eur. J. 2015, 21, 885–892. [Google Scholar] [CrossRef] [PubMed]
  32. Shen, Q.; Li, L.; Li, J.; Tian, H.; Hao, Z. A study on N2O catalytic decomposition over Co/MgO catalysts. J. Hazard. Mater. 2009, 163, 1332–1337. [Google Scholar] [CrossRef] [PubMed]
  33. Luo, J.-Y.; Meng, M.; Li, X.; Li, X.-G.; Zha, Y.-Q.; Hu, T.-D.; Xie, Y.-N.; Zhang, J. Mesoporous Co3O4–CeO2 and Pd/Co3O4–CeO2 catalysts: Synthesis, characterization and mechanistic study of their catalytic properties for low-temperature CO oxidation. J. Catal. 2008, 254, 310–324. [Google Scholar] [CrossRef]
  34. Wang, H.; Ye, J.L.; Liu, Y.; Li, Y.D.; Qin, Y.N. Steam reforming of ethanol over Co3O4/CeO2 catalysts prepared by different methods. Catal. Today 2007, 129, 305–312. [Google Scholar] [CrossRef]
  35. Qiu, N.; Zhang, J.; Wu, Z. Peculiar surface-interface properties of nanocrystalline ceria-cobalt oxides with enhanced oxygen storage capacity. Phys. Chem. Chem. Phys. 2014, 16, 22659–22664. [Google Scholar] [CrossRef] [PubMed]
  36. Vinod, C.P. Surface science as a tool for probing nanocatalysis phenomena. Catal. Today 2010, 154, 113–117. [Google Scholar] [CrossRef]
  37. Hu, P.; Huang, Z.; Amghouz, Z.; Makkee, M.; Xu, F.; Kapteijn, F.; Dikhtiarenko, A.; Chen, Y.; Gu, X.; Tang, X. Electronic Metal-Support Interactions in Single-Atom Catalysts. Angew. Chem. 2014, 126, 3486–3489. [Google Scholar] [CrossRef]
  38. Elias, J.S.; Risch, M.; Giordano, L.; Mansour, A.N.; Shao-Horn, Y. Structure, Bonding, and Catalytic Activity of Monodisperse, Transition-Metal-Substituted CeO2 Nanoparticles. J. Am. Chem. Soc. 2014, 136, 17193–17200. [Google Scholar] [CrossRef] [PubMed]
  39. Cargnello, M.; Fornasiero, P.; Gorte, R.J. Opportunities for tailoring catalytic properties through metal-support interactions. Catal. Lett. 2012, 142, 1043–1048. [Google Scholar] [CrossRef]
  40. Díez-Ramírez, J.; Sánchez, P.; Kyriakou, V.; Zafeiratos, S.; Marnellos, G.E.; Konsolakis, M.; Dorado, F. Effect of support nature on the cobalt-catalyzed CO2 hydrogenation. J. CO2 Util. 2017, 21, 562–571. [Google Scholar] [CrossRef]
  41. Cargnello, M.; Doan-Nguyen, V.V.T.; Gordon, T.R.; Diaz, R.E.; Stach, E.A.; Gorte, R.J.; Fornasiero, P.; Murray, C.B. Control of Metal Nanocrystal Size Reveals Metal-Support Interface Role for Ceria Catalysts. Science 2013, 341, 771–773. [Google Scholar] [CrossRef] [PubMed]
  42. Vayssilov, G.N.; Lykhach, Y.; Migani, A.; Staudt, T.; Petrova, G.P.; Tsud, N.; Skála, T.; Bruix, A.; Illas, F.; Prince, K.C.; et al. Support nanostructure boosts oxygen transfer to catalytically active platinum nanoparticles. Nat. Mater. 2011, 10, 310–315. [Google Scholar] [CrossRef] [PubMed]
  43. Mahammadunnisa, S.K.; Akanksha, T.; Krushnamurty, K.; Subrahmanyam, C. Catalytic decomposition of N2O over CeO2 supported Co3O4 catalysts. J. Chem. Sci. 2016, 128, 1795–1804. [Google Scholar] [CrossRef]
  44. Lin, B.; Liu, Y.; Heng, L.; Ni, J.; Lin, J.; Jiang, L. Effect of ceria morphology on the catalytic activity of Co/CeO2 catalyst for ammonia synthesis. Catal. Commun. 2017, 101, 15–19. [Google Scholar] [CrossRef]
  45. Zabilskiy, M.; Djinović, P.; Tchernychova, E.; Tkachenko, O.P.; Kustov, L.M.; Pintar, A. Nanoshaped CuO/CeO2 Materials: Effect of the Exposed Ceria Surfaces on Catalytic Activity in N2O Decomposition Reaction. ACS Catal. 2015, 5, 5357–5365. [Google Scholar] [CrossRef]
  46. Andrade-Martínez, J.; Ortega-Zarzosa, G.; Gómez-Cortés, A.; Rodríguez-González, V. N2O catalytic reduction over different porous SiO2 materials functionalized with copper. Powder Technol. 2015, 274, 305–312. [Google Scholar] [CrossRef]
  47. Mei, J.; Ke, Y.; Yu, Z.; Hu, X.; Qu, Z.; Yan, N. Morphology-dependent properties of Co3O4/CeO2 catalysts for low temperature dibromomethane (CH2Br2) oxidation. Chem. Eng. J. 2017, 320, 124–134. [Google Scholar] [CrossRef]
  48. Hu, F.; Peng, Y.; Chen, J.; Liu, S.; Song, H.; Li, J. Low content of CoOx supported on nanocrystalline CeO2 for toluene combustion: The importance of interfaces between active sites and supports. Appl. Catal. B Environ. 2019, 240, 329–336. [Google Scholar] [CrossRef]
  49. Lykaki, M.; Pachatouridou, E.; Carabineiro, S.A.C.; Iliopoulou, E.; Andriopoulou, C.; Kallithrakas-Kontos, N.; Boghosian, S.; Konsolakis, M. Ceria nanoparticles shape effects on the structural defects and surface chemistry: Implications in CO oxidation by Cu/CeO2 catalysts. Appl. Catal. B Environ. 2018, 230, 18–28. [Google Scholar] [CrossRef]
  50. Allwar, A.; Md Noor, A.; Nawi, M. Textural Characteristics of Activated Carbons Prepared from Oil Palm Shells Activated with ZnCl2 and Pyrolysis Under Nitrogen and Carbon Dioxide. J. Phys. Sci. 2008, 19, 93–104. [Google Scholar]
  51. Kumar, S.; Sharma, C. Synthesis, characterization and catalytic wet air oxidation property of mesoporous Ce1-xFexO2 mixed oxides. Mater. Chem. Phys. 2015, 155, 223–231. [Google Scholar]
  52. Tan, L.; Tao, Q.; Gao, H.; Li, J.; Jia, D.; Yang, M. Preparation and catalytic performance of mesoporous ceria-base composites CuO/CeO2, Fe2O3/CeO2 and La2O3/CeO2. J. Porous Mater. 2017, 24, 795–803. [Google Scholar] [CrossRef]
  53. Farahmandjou, M.; Zarinkamar, M. Synthesis of nano-sized ceria (CeO2) particles via a cerium hydroxy carbonate precursor and the effect of reaction temperature on particle morphology. J. Ultrafine Grained Nanostructured Mater. 2015, 48, 5–10. [Google Scholar]
  54. Sharma, V.; Eberhardt, K.M.; Sharma, R.; Adams, J.B.; Crozier, P.A. A spray drying system for synthesis of rare-earth doped cerium oxide nanoparticles. Chem. Phys. Lett. 2010, 495, 280–286. [Google Scholar] [CrossRef]
  55. Liu, J.; Zhao, Z.; Wang, J.; Xu, C.; Duan, A.; Jiang, G.; Yang, Q. The highly active catalysts of nanometric CeO2-supported cobalt oxides for soot combustion. Appl. Catal. B Environ. 2008, 84, 185–195. [Google Scholar] [CrossRef]
  56. Yu, S.W.; Huang, H.H.; Tang, C.W.; Wang, C.B. The effect of accessible oxygen over Co3O4-CeO2 catalysts on the steam reforming of ethanol. Int. J. Hydrogen Energy 2014, 39, 20700–20711. [Google Scholar] [CrossRef]
  57. Konsolakis, M.; Carabineiro, S.A.C.; Marnellos, G.E.; Asad, M.F.; Soares, O.S.G.P.; Pereira, M.F.R.; Órfão, J.J.M.; Figueiredo, J.L. Effect of cobalt loading on the solid state properties and ethyl acetate oxidation performance of cobalt-cerium mixed oxides. J. Colloid Interface Sci. 2017, 496, 141–149. [Google Scholar] [CrossRef] [PubMed]
  58. Mock, S.A.; Sharp, S.E.; Stoner, T.R.; Radetic, M.J.; Zell, E.T.; Wang, R. CeO2 nanorods-supported transition metal catalysts for CO oxidation. J. Colloid Interface Sci. 2016, 466, 261–267. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, Y.; Hu, X.; Zheng, K.; Zhang, H.; Zhao, Y. Effect of precipitants on the catalytic activity of Co–Ce composite oxide for N2O catalytic decomposition. Reac. Kinet. Mech. Cat. 2018, 123, 707–721. [Google Scholar] [CrossRef]
  60. Wang, L.; Liu, H. Mesoporous Co-CeO2 catalyst prepared by colloidal solution combustion method for reverse water-gas shift reaction. Catal. Today 2018, 316, 155–161. [Google Scholar] [CrossRef]
  61. Kumar Megarajan, S.; Rayalu, S.; Teraoka, Y.; Labhsetwar, N. High NO oxidation catalytic activity on non-noble metal based cobalt-ceria catalyst for diesel soot oxidation. J. Mol. Catal. A Chem. 2014, 385, 112–118. [Google Scholar] [CrossRef]
  62. Konsolakis, M. The role of Copper–Ceria interactions in catalysis science: Recent theoretical and experimental advances. Appl. Catal. B Environ. 2016, 198, 49–66. [Google Scholar] [CrossRef]
  63. Cui, Y.; Dai, W.-L. Support morphology and crystal plane effect of Cu/CeO2 nanomaterial on the physicochemical and catalytic properties for carbonate hydrogenation. Catal. Sci. Technol. 2016, 6, 7752–7762. [Google Scholar] [CrossRef]
  64. Guo, X.; Zhou, R. A new insight into the morphology effect of ceria on CuO/CeO2 catalysts for CO selective oxidation in hydrogen-rich gas. Catal. Sci. Technol. 2016, 6, 3862–3871. [Google Scholar] [CrossRef]
  65. Wang, C.; Cheng, Q.; Wang, X.; Ma, K.; Bai, X.; Tan, S.; Tian, Y.; Ding, T.; Zheng, L.; Zhang, J.; et al. Enhanced catalytic performance for CO preferential oxidation over CuO catalysts supported on highly defective CeO2 nanocrystals. Appl. Surf. Sci. 2017, 422, 932–943. [Google Scholar] [CrossRef]
  66. Dou, J.; Tang, Y.; Nie, L.; Andolina, C.M.; Zhang, X.; House, S.; Li, Y.; Yang, J.; Tao, F.F. Complete Oxidation of Methane on Co3O4/CeO2 Nanocomposite: A Synergic Effect. Catal. Today 2018, 311, 48–55. [Google Scholar] [CrossRef]
  67. Konsolakis, M.; Sgourakis, M.; Carabineiro, S.A.C. Surface and redox properties of cobalt-ceria binary oxides: On the effect of Co content and pretreatment conditions. Appl. Surf. Sci. 2015, 341, 48–54. [Google Scholar] [CrossRef]
  68. Ma, J.; Rodriguez, N.M.; Vannice, M.A.; Baker, R.T.K. Nitrous oxide decomposition and reduction over copper catalysts supported on various types of carbonaceous materials. Top. Catal. 2000, 10, 27–38. [Google Scholar] [CrossRef]
  69. You, Y.; Chang, H.; Ma, L.; Guo, L.; Qin, X.; Li, J.; Li, J. Enhancement of N2O decomposition performance by N2O pretreatment over Ce-Co-O catalyst. Chem. Eng. J. 2018, 347, 184–192. [Google Scholar] [CrossRef]
  70. Iwanek, E.; Krawczyk, K.; Petryk, J.; Sobczak, J.W.; Kaszkur, Z. Direct nitrous oxide decomposition with CoOx-CeO2 catalysts. Appl. Catal. B Environ. 2011, 106, 416–422. [Google Scholar] [CrossRef]
  71. Piskorz, W.; Zasada, F.; Stelmachowski, P.; Kotarba, A.; Sojka, Z. Decomposition of N2O over the surface of cobalt spinel: A DFT account of reactivity experiments. Catal. Today 2008, 137, 418–422. [Google Scholar] [CrossRef]
  72. Abu-Zied, B.M.; Soliman, S.A.; Abdellah, S.E. Pure and Ni-substituted Co3O4 spinel catalysts for direct N2O decomposition. Chin. J. Catal. 2014, 35, 1105–1112. [Google Scholar] [CrossRef]
  73. Stelmachowski, P.; Maniak, G.; Kotarba, A.; Sojka, Z. Strong electronic promotion of Co3O4 towards N2O decomposition by surface alkali dopants. Catal. Commun. 2009, 10, 1062–1065. [Google Scholar] [CrossRef]
  74. Savereide, L.; Nauert, S.L.; Roberts, C.A.; Notestein, J.M. The effect of support morphology on CoOX/CeO2 catalysts for the reduction of NO by CO. J. Catal. 2018, 366, 150–158. [Google Scholar] [CrossRef]
  75. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  76. Barthos, R.; Hegyessy, A.; Klébert, S.; Valyon, J. Vanadium dispersion and catalytic activity of Pd/VOx/SBA-15 catalysts in the Wacker oxidation of ethylene. Microporous Mesoporous Mater. 2015, 207, 1–8. [Google Scholar] [CrossRef]
  77. Xu, J.; Harmer, J.; Li, G.; Chapman, T.; Collier, P.; Longworth, S.; Tsang, S.C. Size dependent oxygen buffering capacity of ceria nanocrystals. Chem. Commun. 2010, 46, 1887–1889. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The BJH (Barret-Joyner-Halenda) desorption pore size distribution (PSD) and (b) the adsorption–desorption isotherms of CeO2 and Co/CeO2 samples.
Figure 1. (a) The BJH (Barret-Joyner-Halenda) desorption pore size distribution (PSD) and (b) the adsorption–desorption isotherms of CeO2 and Co/CeO2 samples.
Catalysts 09 00233 g001
Figure 2. The XRD patterns of the CeO2 and Co/CeO2 samples.
Figure 2. The XRD patterns of the CeO2 and Co/CeO2 samples.
Catalysts 09 00233 g002
Figure 3. The transmission electron microscopy images of CeO2 (ac) and Co/CeO2 (df) samples: (a) CeO2-NR, (b) CeO2-NP, (c) CeO2-NC, (d) Co/CeO2-NR, (e) Co/CeO2-NP and (f) Co/CeO2-NC.
Figure 3. The transmission electron microscopy images of CeO2 (ac) and Co/CeO2 (df) samples: (a) CeO2-NR, (b) CeO2-NP, (c) CeO2-NC, (d) Co/CeO2-NR, (e) Co/CeO2-NP and (f) Co/CeO2-NC.
Catalysts 09 00233 g003
Figure 4. The H2-TPR profiles of (a) bare CeO2 and (b) Co3O4, Co/CeO2 samples.
Figure 4. The H2-TPR profiles of (a) bare CeO2 and (b) Co3O4, Co/CeO2 samples.
Catalysts 09 00233 g004
Figure 5. The X-ray photoelectron spectroscopy (XPS) spectra of (a) Ce 3d and (b) O 1s of bare CeO2 and Co/CeO2 samples.
Figure 5. The X-ray photoelectron spectroscopy (XPS) spectra of (a) Ce 3d and (b) O 1s of bare CeO2 and Co/CeO2 samples.
Catalysts 09 00233 g005
Figure 6. The Co 2p XPS spectra of the Co3O4 and Co/CeO2 samples: The Co 2p XPS spectra of Co/CeO2 samples have been magnified.
Figure 6. The Co 2p XPS spectra of the Co3O4 and Co/CeO2 samples: The Co 2p XPS spectra of Co/CeO2 samples have been magnified.
Catalysts 09 00233 g006
Figure 7. N2O conversion as a function of temperature for CeO2 and Co/CeO2 samples of different morphology (a) in the absence and (b) in the presence of oxygen: The reaction conditions are 1000 ppm N2O, 0 or 2% O2 and Gas Hour Space Velocity (GHSV) = 40,000 h−1.
Figure 7. N2O conversion as a function of temperature for CeO2 and Co/CeO2 samples of different morphology (a) in the absence and (b) in the presence of oxygen: The reaction conditions are 1000 ppm N2O, 0 or 2% O2 and Gas Hour Space Velocity (GHSV) = 40,000 h−1.
Catalysts 09 00233 g007
Figure 8. The half-conversion temperature (T50) as a function of the TPR surface-to-bulk oxygen ratio (Os/Ob).
Figure 8. The half-conversion temperature (T50) as a function of the TPR surface-to-bulk oxygen ratio (Os/Ob).
Catalysts 09 00233 g008
Table 1. The textural and structural properties of bare CeO2 and Co/CeO2 samples.
Table 1. The textural and structural properties of bare CeO2 and Co/CeO2 samples.
SampleBET AnalysisXRD Analysis
BET Surface Area (m2 g−1)Pore Volume (cm3/g)Average Pore Diameter (nm)Crystallite Size (nm), DXRD 1
CeO2Co3O4
CeO2-NC370.2627.427 ± 1-
CeO2-NR790.4824.215 ± 1-
CeO2-NP880.177.911 ± 1-
Co/CeO2-NC280.1522.624 ± 119 ± 1
Co/CeO2-NR720.3117.414 ± 116 ± 1
Co/CeO2-NP710.179.811 ± 115 ± 1
1 Calculated applying the Williamson–Hall plot after the Rietveld refinement of diffractograms.
Table 2. The redox properties of the bare CeO2 and Co/CeO2 samples.
Table 2. The redox properties of the bare CeO2 and Co/CeO2 samples.
SampleH2 Consumption (mmol H2 g−1) aOs/Ob RatioPeak Temperature (°C)
Os PeakOb PeakTotalOs PeakOb Peak
CeO2-NP0.480.510.990.94555804
CeO2-NR0.590.521.111.13545788
CeO2-NC0.410.580.990.71589809
Peaks a+bCeO2 PeakTotal Peak aPeak b
Co/CeO2-NP2.400.613.01 333388
Co/CeO2-NR2.370.622.99 318388
Co/CeO2-NC2.050.322.37 335405
a Estimated by the area of the corresponding temperature programmed reduction (TPR) peaks, which is calibrated against a known amount of CuO standard sample.
Table 3. The XPS results of bare CeO2 and Co/CeO2 samples.
Table 3. The XPS results of bare CeO2 and Co/CeO2 samples.
SampleCo2+/Co3+Ce3+ (%)OII/OI
CeO2-NC-23.30.50
CeO2-NR-24.30.47
CeO2-NP-25.30.49
Co/CeO2-NC1.0626.10.51
Co/CeO2-NR1.3228.50.60
Co/CeO2-NP0.9426.70.53
Table 4. The N2O conversion and specific activity of Co/CeO2 samples at 420 °C: The reaction conditions are 1000 ppm N2O, 0 or 2 vol. % O2 and GHSV = 40,000 h−1.
Table 4. The N2O conversion and specific activity of Co/CeO2 samples at 420 °C: The reaction conditions are 1000 ppm N2O, 0 or 2 vol. % O2 and GHSV = 40,000 h−1.
SampleN2O Conversion (%)Specific Activity
O2 AbsenceO2 PresenceO2 AbsenceO2 Presence
r (nmol g−1 s−1)r (nmol m−2 s−1)r (nmol g−1 s−1)r (nmol m−2 s−1)
Co/CeO2-NC16.28.61665.9883.1
Co/CeO2-NP20.210.72072.91091.5
Co/CeO2-NR25142563.61432.0

Share and Cite

MDPI and ACS Style

Lykaki, M.; Papista, E.; Kaklidis, N.; Carabineiro, S.A.C.; Konsolakis, M. Ceria Nanoparticles’ Morphological Effects on the N2O Decomposition Performance of Co3O4/CeO2 Mixed Oxides. Catalysts 2019, 9, 233. https://doi.org/10.3390/catal9030233

AMA Style

Lykaki M, Papista E, Kaklidis N, Carabineiro SAC, Konsolakis M. Ceria Nanoparticles’ Morphological Effects on the N2O Decomposition Performance of Co3O4/CeO2 Mixed Oxides. Catalysts. 2019; 9(3):233. https://doi.org/10.3390/catal9030233

Chicago/Turabian Style

Lykaki, Maria, Eleni Papista, Nikolaos Kaklidis, Sόnia A. C. Carabineiro, and Michalis Konsolakis. 2019. "Ceria Nanoparticles’ Morphological Effects on the N2O Decomposition Performance of Co3O4/CeO2 Mixed Oxides" Catalysts 9, no. 3: 233. https://doi.org/10.3390/catal9030233

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop